• Aucun résultat trouvé

The Convergence of Systems and Reductionist Approaches in Complex Trait Analysis.

N/A
N/A
Protected

Academic year: 2021

Partager "The Convergence of Systems and Reductionist Approaches in Complex Trait Analysis."

Copied!
10
0
0

Texte intégral

(1)

Leading Edge

Perspective

The Convergence of Systems and

Reductionist Approaches in Complex Trait Analysis

Evan G. Williams1and Johan Auwerx1,*

1Laboratory of Integrative and Systems Physiology, E´cole Polytechnique Fe´de´rale de Lausanne, 1015 Lausanne, Switzerland *Correspondence:admin.auwerx@epfl.ch

http://dx.doi.org/10.1016/j.cell.2015.06.024

Research into the genetic and environmental factors behind complex trait variation has traditionally

been segregated into distinct scientific camps. The reductionist approach aims to decrypt

pheno-typic variability bit by bit, founded on the underlying hypothesis that genome-to-phenome relations

are largely constructed from the additive effects of their molecular players. In contrast, the systems

approach aims to examine large-scale interactions of many components simultaneously, on the

premise that interactions in gene networks can be both linear and non-linear. Both approaches

are complementary, and they are becoming increasingly intertwined due to developments in

gene editing tools, omics technologies, and population resources. Together, these strategies are

beginning to drive the next era in complex trait research, paving the way to improve agriculture

and toward more personalized medicine.

Over the past century, great strides have been made toward identifying and understanding the genetic and environmental factors affecting complex traits. This progress has come from divergent though complementary genetic techniques that address similar biological hypotheses from different an-gles: the study of humans and crops versus model organisms, forward versus reverse genetics, holistic versus additive models, and so forth. Results from each of these approaches include the identification of thousands of variants influencing complex traits (Kingsmore et al., 2007) and the mechanistic delineation of hundreds of molecular pathways (Kanehisa et al., 2014). This knowledge has led to numerous practical benefits, ranging from the development of vaccines for trans-missible diseases to drug treatments, dietary, and lifestyle changes for metabolic diseases to the rational modification of crops and livestock for agricultural needs. However, our ability to predict when, where, and how genetic, environ-mental, or gene-by-environment interactions (G3 E) will lead to specific phenotypic outcomes is still limited, and few ratio-nally generated cures or preventative strategies are available for common diseases such as cancer and diabetes. In large part, this shortcoming stems from the challenge of identifying key genetic and environmental regulators that can be targeted by drugs, lifestyle changes, or genetic modifications. For de-cades, the identification of new targets and the implementa-tion of treatments have been hampered by the conceptual complexity and diversity of genetic mechanisms. This obstacle is now being overcome by a combination of improvements in technical capability, new approaches in scientific thought, and increased resource sharing. With this in mind, how can we use these developments to find new medical and biological breakthroughs? To answer this, we must first consider the current state of research into complex traits and how it has evolved over time (Figure 1A).

Developments in Complex Trait Analysis

(2)

trait variation in natural populations, which are generally more subtle (Chakravarti et al., 2013; MacArthur et al., 2012). Minor variants, gene3 gene, and G 3 E interactions can be examined mechanistically using modern G/LOF tools, yet the exponential increase in the number of such possibilities as complexity ex-pands necessitates the use of prior hypotheses instead of unbiased screens, particularly for vertebrate research. Finally, mechanisms that are uncovered in G/LOF models may not necessarily be generalizable to natural populations, whether in humans or in agriculture. These limitations of G/LOF models were recognized from the outset (Capecchi, 2005), but potential alternatives—particularly, population genetics—suffered from strong deficits as well.

In parallel to the developments in forward- and reverse-ge-netics techniques, progress continued steadily on molecular measurement technologies that expanded the scope and depth of genetic analysis. The debut of what has become the ‘‘omics revolution’’ began with massive investments in large-scale nucleotide sequencing (Smith and Hood, 1987), biological appli-cations of mass spectrometry (Fenn et al., 1989; Wasinger et al., 1995), and array technology (Schena et al., 1995). By the late 1990s, the genomic and transcriptomic tools were sufficiently refined and affordable that small collaborative groups had the capability to generate and test hypotheses that required full pathway analysis by using comprehensive genomic and

tran-Figure 1. Broad Summary of Concepts and Developments in Genetic Analysis

(A) Relative timeline of complex trait analysis, listing some key landmarks and their approximate dates.

(B) (Left) The forward-genetics approach relies on identifying divergent phenotypes and then searching for the causative genetic factors, generally through QTL analysis or GWAS. (Right) The reverse-genetics approach relies on modifying a gene (or genes) of interest and then scanning for impact on downstream traits, often by G/LOF modifications.

(C) Single genes can fully regulate a phenotype (e.g., sickle-cell trait), or they can play a role in more complex phenotypes for part of the popula-tion (e.g., PPARg and metabolic disease;Deeb et al., 1998). However, complex trait variation also comes from epistasis (e.g., mutations in p53 and other genes together are necessary to develop cancer;Soussi et al., 1994). To fully explain heri-tability and complex trait variation for an entire population (e.g., diabetes), it is likely that we will need the ability to model and test arbitrarily com-plex interaction of dozens or more variants simul-taneously. In this cartoon, examples are given in which one, two, or many genes must be modified in order to change the observed phenotype (color, monogenic; spots, oligogenic; or size, polygenic).

scriptomic datasets. While the resulting and unprecedentedly thorough data sets aided both population and G/LOF re-search, they particularly boosted the population approach. In theory, omics coverage could provide the capacity to identify causal gene networks wholesale through data-driven approaches—even directly in humans. Indeed, initial results using this approach to study common com-plex disorders were promising, as exemplified by the identifica-tion of variants in two genes, PPARg (Deeb et al., 1998) and MC4R (Yeo et al., 1998), causal for metabolic disease. However, human population studies examining such genome-to-phenome links (e.g., genome-wide association studies [GWAS]) ran into several major barriers, among them the issues of linkage disequi-librium, commonly detected SNPs having small effect sizes, poor long-term environmental control, and the perennial issue of ‘‘missing heritability’’ (Goldstein, 2009; Lander, 2011). Further-more, while genotype information is fairly consistent across time and tissue, the ephemeral nature of transcripts, proteins, and metabolites hindered detailed mechanistic analyses in human populations due to the difficulty or impossibility—depending on tissue—of obtaining biopsies. In retrospect, we have now seen that human GWAS led only to a slow trickle of discoveries be-tween novel gene variants and complex traits (McCarthy et al., 2008).

(3)

genetic reference populations (GRPs). However, the imple-mentation of these concepts during the early years of high-throughput sequencing and microarray transcriptomics was hindered by small cohort sizes and the limited selection of ready-made GRPs (Williams et al., 2001). The earliest GRPs, such as the BXD mice (Taylor et al., 1973), had been developed and utilized decades previously, primarily for forward-genetics analyses. However, such early studies often linked phenotypic variants to broad quantitative trait loci (QTLs) containing dozens to hundreds of candidate genes; thus, the specific causative genes were rarely identified (Flint et al., 2005). This shortcoming could be largely solved by analyzing larger populations and by improving recombination density (Darvasi and Soller, 1995), yet the development of GRPs is expensive and can take many years to generate, particularly in vertebrates. Thus, little development was done on expanding model organism populations—even in invertebrates—until technological advances in omics and the shortcomings (and benefits) of human population research had been well-recognized. Since then, a diverse collection of GRPs

Figure 2. Model Population Development and Analysis

(A) A summary and basic breeding schema for major population types. F2s are generated ad hoc, which can be studied or crossed and inbred for 20+ generations to generate recombinant inbred lines (RIL). F2s can be backcrossed to generate congenic, consomic, or conplastic strains. (B) Heritability should be considered as a function of gene-environment interactions, especially for complex traits influenced by environmental fac-tors. (Upper-left) Across a diverse GRP, the observed trait is significantly elevated in ‘‘Envi-ronment B.’’ (Lower-left) Despite the difference in trait expression across environments, the herita-bility is similar when groups are segregated, though it plummets when the groups are com-bined. If the environmental difference can be controlled and groups separated, as here, the environment and G3 E factors can be calculated by two-way ANOVA. (Right) When the same data are displayed as strain averages, the heritability drop can be visualized. Within either environment, variability due to genotype (y axis) is far lower than the error within a genotype (error bars). When these data are compressed without respect to environment, the cross-genotype variance compared to within-genotype variance decreases, hence the drop in observed heritability.

(C) A broad overview of relative species particu-larities and strengths for the most common model organisms in population genetics. The suitability of any model also depends substantially on the question to be addressed and specific experi-mental design.

has been developed across many model organisms that can simulate many as-pects of the genetic complexity of natural populations (human, animal, or plant) but in controlled settings (Churchill et al., 2004; Kover et al., 2009; Mackay et al., 2012) (Table 1).

(4)

that in turn display gene expression from very low (akin to ‘‘knockouts’’) to average (akin to ‘‘wildtype’’) to very high (akin to ‘‘transgenic overexpression’’)—and everywhere in between (Albert et al., 2014). In such extensive populations, nearly all genes display very strong levels of variability. Moreover, the po-wer to detect QTLs depends largely on the variance of the traits of interest. Thus, this approach (dubbed extreme QTL, or ‘‘X-QTL,’’ mapping) increases both the scale and scope of find-ings as compared using smaller cohorts. However, two reasons limit the generation of X-QTLs in more complex organisms. The first is due to scale, as analyzing millions of individuals remains logistically and economically prohibitive in most organisms, and the second due to technique, as the expression of any gene can be examined visually in yeast (and C. elegans) by tagging target genes with green fluorescent protein (GFP). As such, gene expression may be assessed across millions of indi-viduals visually, groups may be separated into a few clusters based on expression, and omics analysis can be performed only on these subsets. For more complex organisms, such as

mice, omics analyses would be necessary for all millions of indi-viduals to capture the full spectrum of variance.

Model populations can also address many of the shortcomings of studies using G/LOF models or single inbred lines. For in-stance, early drug trials in model organisms are predominantly performed as experiments with a genotypic n = 1. The reality of complex trait genetics, however, means that even a reproducible finding in one individual may not apply in another. For example, had researchers exclusively used the DBA/2J mouse strain in place of the prototypical C57BL/6J, they would have concluded that morphine is a non-addictive and inefficient painkiller (Elmer et al., 2010). Examples like this have led to a recognition that genetic diversity must be considered at all steps from target iden-tification to final outcome, an approach that is beginning to be implemented medically (Barretina et al., 2012) and that is starting to assist in the design of compounds that are highly effective in particular, defined subsets of patients (e.g., CFTR variant-specific drug treatments for cystic fibrosis; Ledford, 2012). However, simplified models remain essential: early drug screens of dozens

Table 1. Landmark Dates and Papers in the Generation of Model Populations, along with Key Populations or First Populations in a Variety of Species

Date Resource Organism Reference

Wild-Type (All)

1866 F2 intercross Pea G.J. Mendel

1909 Inbred lines Mouse C.C. Little

1959 Recombinant inbred (CXB) Mouse (Bailey, 1981)

1971 Recombinant inbred (BXD) Mouse (Taylor et al., 1973)

1982 Recombinant inbred (BXH/HXB) Rat (Pravenec et al., 1989)

1984 Diversity panel (N/Nih) Rat (Li and Lumeng, 1984)

1988 Recombinant inbred (WSxM13) Corn (Burr et al., 1988)

1996 Recombinant inbred (WSxM13) Arabidopsis (Liu et al., 1996)

1997 Recombinant inbred (Rx2b) Drosophila (Nuzhdin et al., 1997)

2000 Chromosome substitution Mice (Nadeau et al., 2000)

2001 Recombinant inbred (BOxRC301) Caenorhabditis (Ayyadevara et al., 2001)

2002 Recombinant inbred Soybean (Yuan et al., 2002)

2003 Recombinant inbred Barley (Arru et al., 2003)

2004 Advanced recombinant inbred (BXD) Mouse (Peirce et al., 2004)

2004 Recombinant inbred (LXS) Mouse (Williams et al., 2004)

2004 Advanced recombinant inbred (collaborative cross) Mouse (Churchill et al., 2004) 2005 Diversity panel (heterogeneous stock) Caenorhabditis (Sivasundar and Hey, 2005)

2006 Diversity panel (heterogeneous stock) Mouse (Valdar et al., 2006)

2008 Diversity panel (1001 Genomes Project) Arabidopsis (Ossowski et al., 2008) 2009 Advanced recombinant inbred (MAGIC) Arabidopsis (Kover et al., 2009)

2009 Advanced recombinant inbred (N2xCB4856) Caenorhabditis (Rockman and Kruglyak, 2009)

2010 Chromosome substitution Rice (Xu et al., 2010)

2010 Diversity panel (hybrid mouse diversity panel) Mouse (Bennett et al., 2010)

2011 Recombinant inbred (four-way cross) Yeast (Cubillos et al., 2011)

2012 Diversity panel (DGRP) Drosophila (Mackay et al., 2012)

2012 Advanced recombinant inbred (DSPR) Drosophila (King et al., 2012)

(5)

of novel compounds to treat an illness cannot be efficiently per-formed if they are each thoroughly tested in extensive GRPs. These tradeoffs between comprehensive genetic models (which may be excessively complex) and simplified models (which may be oversimplified) must be considered at each stage of the research process. This realization is beginning to lead to the com-plementary implementation of these approaches throughout the pipeline of complex trait analysis.

Complementary Approaches in Complex Trait Genetics The parallel expansion of gene editing capabilities, omics tech-nologies, and population resources has created the capability to implement experiments incorporating mechanistic in vitro experiments, comprehensive studies in G/LOF organisms and model populations, and research in natural populations. This complementarity is multidirectional. In one way, in vitro mecha-nistic studies can provide detailed hypotheses for in vivo verifi-cation using reductionist models, then tested in diverse genetic backgrounds for potential interaction effects of the variant, and finally examined outside of the laboratory setting. Conversely, population studies—often beset by false discovery—may use G/LOF models to distinguish several potential causal factors or to obtain detailed mechanistic understanding. In general, study design can also be streamlined by considering complementary cross-species techniques. For instance, the de novo identifica-tion of gene-phenotype links can be readily established using

un-Figure 3. A Summary of a Few Study De-signs Using Model and/or Cross-Species Approaches

(A) Conceptual schematic of G/LOF screenings performed in simpler model organisms (e.g.,

Drosophila), which may be used to generate

tar-geted hypotheses for G/LOF studies in more complex (e.g., mammalian) models. These results may then provide targeted hypotheses for valida-tion in human GWAS.

(B) Cross-model example of population studies benefitting from one-another. QTL results do not indicate a specific gene, and equivalent GWAS results, displayed as a Manhattan plot, would not approach significance on a genome-wide scan. Furthermore, candidate SNPs in GWAS may be in linkage disequilibrium. The complementarity of independent population studies can often be applied across species by comparing syntenic regions.

(C) Conceptual approach of gene transfer. Natural genetic pathways that allow certain plants to resist stresses such as drought or disease may poten-tially be translated into other species via trans-genesis.

biased screens in invertebrates, thus providing specific hypotheses for G/LOF work in mammals—and eventually for the clinic (Figure 3A). Population studies as well may benefit from cross-species analysis—from GRP to GWAS or vice versa—such as in assisting the identifica-tion of causal genes under a QTL or within linkage disequilibrium (Figure 3B). Finally, it is not fundamentally necessary for the target genes, or even mechanisms, to be conserved across species to derive utility from a cross-species or cross-model approach. Discovery of unique aspects of an organism can be as informative as the commonalities, such as the key differences in atherosclerosis development between mice and humans (Jiang et al., 1992). Furthermore, particularly for agriculture, the identification of species-specific genes and mechanisms can also assist in trans-genic development of crops, which are resistant to particular diseases or environmental conditions (Figure 3C).

(6)

separate recent example comparing results from a genetic loss-of-function screen in Drosophila to rare diseases in human exome data (Yamamoto et al., 2014). The complementarity of these cross-species exchanges can be quite surprising; for instance, genes linked to neurological diseases in humans can be effectively analyzed in reductive plant models (Xu and Møller, 2011).

In other cases, both hypothesis generation and validation may be performed using population studies, including for humans and mice (related toFigure 3B). For instance, a recent extensive metabolic study of the BXD mouse GRP identified more than a dozen novel and significant phenotype QTLs (Andreux et al., 2012). The QTL for systolic blood pressure, with five candidate genes, was prioritized for validation due to the small number of possibilities and, moreover, the ready availability of independent population studies that measured blood pressure (Koutnikova et al., 2009). In the subsequent analysis of three independent hu-man GWAS, SNPs in a single gene, Ubp1, were associated consistently and significantly with elevated blood pressure. Although these SNPs were nominally significant in the original data, the correction for multiple testing across hundreds of thou-sands of SNPs across the whole genome drowned out the signal. Details about cellular mechanisms can also be confidently attrib-uted solely using population data. In a separate BXD study, it was observed that diabetic mice have significantly lower levels of the metabolite 2-aminoadipate, independently from the ef-fects of low- or high-fat diet feeding (Wu et al., 2014). In turn, the levels of this metabolite mapped significantly to the gene lo-cus containing Dhtkd1, a known enzyme in 2-aminoadipate metabolism (Danhauser et al., 2012). Thus, a mechanistic link was identified between Dhtkd1 and diabetes via the regulation of 2-aminoadipate. Furthermore, two independent populations (one mouse, one human) were examined using this hypothesis, and again the same links with diabetes were observed. As before, multiple testing correction had prevented the identifica-tion of this connecidentifica-tion in the original data sets (Wu et al., 2014). Finally, human medical application is not the only goal: dis-eases uniquely affecting animals (e.g., rinderpest, foot-and-mouth disease) and plants (e.g., coffee rust, black rot) are equally important. Additionally, genetic engineering in agriculture in response to environmental needs or changes (e.g., drought resistance, improved yield) is a critical research goal, as we must contend with expanding populations and general climatic variation. For instance, gene-phenotype links identified in Arabi-dopsis—itself not a crop—are often applied to agriculture. Muta-genesis screens in Arabidopsis have identified genetic variants influencing flowering promotion (e.g., SFT) and flowering repres-sion (e.g., SP), which can be subsequently applied cross-spe-cies (related toFigure 3C). In tomato, the orthologs of these genes were rationally modified, leading to a striking 130% varia-tion in tomato yield between the suppressed and promoted plants (Park et al., 2014). Similar cross-species plant studies have improved cold tolerance in tobacco (Zhao et al., 2009), drought resistance in rice (Datta et al., 2012), and salt resistance in barley (Schilling et al., 2014). Research for agricultural genetics is often parallel to medical genetics, yet this divergence is not fundamental, and the two can benefit equally from sharing tech-nologies, methodologies, and findings.

Big Data

Moving across species and between or among populations and G/LOF models can provide a greater understanding of biological mechanisms underlying complex traits. However, the in vivo application of such resources still requires extensive experience and potentially time and expense. To this end, it is essential to consider and exploit the wealth of results available in public ‘‘big data’’ resources, which can allow hypothesis examination in silico. Over the last 10 years, the generation of full genomic and transcriptomic data sets has become commonplace. The full DNA sequence of more than 100 species on the UCSC Genome Browser (Kent et al., 2002), more than 1.2 million com-plete transcript data sets on the Gene Expression Omnibus (GEO) (Barrett et al., 2013), and dozens of other smaller reposi-tories have sprung up as well, often organized per species or per topic (Bastian et al., 2008; Chesler et al., 2004; Harris et al., 2010). Conjointly, a wide array of software and web re-sources have been developed to aid in the analysis and interpre-tation of large systems data sets (Barrett et al., 2013; Subrama-nian et al., 2005; Wang et al., 2003). These resources can provide secondary study validation, study power, and founding hypoth-eses and can inform decisions on study design. These possibil-ities are rapidly expanding, as studies increasingly generate more data than can be analyzed under the scope of a single pro-gram. When made public, this ‘‘excess’’ information adds to a treasure trove for secondary analysis and independent valida-tion. Consequently, primary research projects are increasingly taking advantage of independent research beyond the introduc-tion and verbatim citaintroduc-tions—whether for hypothesis generaintroduc-tion, validation, or complete meta-analysis (Khan et al., 2014; Shin et al., 2014).

Historical data can also grow in value over time as additional, diverse data sets are collated. For example, early QTL studies often identified strong gene-phenotype links but were unable to establish the causal genes. These ‘‘unfinished stories’’ now serve as pre-made hypotheses, such as for detailed longevity data recorded in the BXD GRP well before the common imple-mentation of omics technologies (De Haan and Van Zant, 1999; Gelman et al., 1988). These two early studies observed a common, significant QTL on chromosome 2 that contributed about 30% to the overall variation in lifespan, but the causal gene(s) were not defined, and the QTL went unproven. More than a decade later, using modern genetic maps, transcriptomic data, and information on sequence variants, specific candidate genes were found under the QTL. Subsequently, two of these genes (mrps-5 and nkcc-1) were validated in C. elegans using G/LOF technologies that, again, did not exist at the time of the initial study—even if candidate genes had been identified (Houtkooper et al., 2013). Further mechanistic studies on mrps-5 in C. elegans and in an independent set of BXDs were used to uncover that the longevity effects stemmed from the dysregulation of mitochondrial protein translation. The resulting stoichiometric imbalance between mitochondrial and nuclear-encoded proteins, dubbed mitonuclear imbalance, induces the mitochondrial unfolded protein response (known as UPRmt), a

(7)

of detailed electronic medical records at hospitals now serve as the backbone for a recent approach called phenome-wide asso-ciation studies (PheWAS)—essentially, the reverse-genetics analog of GWAS (Denny et al., 2013). In GWAS, a phenotype is queried against genotypes to determine the causative gene(s), whereas in PheWAS, a gene or SNP of interest is queried against phenotypes to determine associated traits. This is conceptually straightforward, yet to be effective, this approach requires extensive and comprehensive sets of phenotype data—a far more expensive and time-consuming task to generate than for genotype data. Some successes have already been reported with this approach, e.g., FTO—traditionally an obesity gene— was recently linked to fibrocystic breast disease by PheWAS (Cronin et al., 2014).

The increasing richness of omics data from both populations and G/LOF models also allows meta-analytical studies to detect and validate novel findings completely in silico (Horvath, 2013; Lee et al., 2012). In one recent high-profile example of meta-analysis, thousands of human methylation arrays spanning 51 tissues were cross-referenced with a single basic phenotype— the age of the patient (Horvath, 2013). A statistical model was then developed to accurately calculate the ages of the patients using only tissue samples and a set of DNA methylation sites. Using this model, it was observed that cancers have ‘‘older’’ methylation patterns than healthy cells; for instance, calculating the age of a 30-year-old breast cancer patient using a blood sample would yield an age of 27–33 years, while the same calcu-lation using a breast tumor biopsy would yield an age of75 years. The confidence in the study results was made possible by the wealth of independently generated methylation data sets: half of the studies were used to generate hypotheses and the other half to test hypotheses, thus providing experimental validation without a need for ‘‘new’’ data. The crossover and ‘‘re-use’’ of data is not new, yet the viability of this approach has rapidly improved due to the remarkable increase in the qual-ity and scope of omics data sets. However, the informatic diffi-culty in large-scale manipulation of this approach has thus far largely restricted its application to computational biology (Marx, 2013). Correspondingly, there has been a disconnect be-tween bioinformatics groups working in silico and generating non-validated networks of profound complexity and traditional in vivo and in vitro biologists elucidating biochemical processes that may be only applicable in the specific circumstances of a single experiment. As with the converging applications of popu-lation genetics and G/LOF models, computational and wet lab biologists are now beginning to find common ground; bioinfor-matics platforms are becoming more widely established both institutionally and in individual wet labs.

Future Perspectives

The recent fundamental shifts we have discussed are beginning to drive the next era in complex trait analysis, yet it is only the next step of many in our push to understand genetics. We have mastered the ability to model and analyze monogenic-gene-to-phenotype interactions, both through their identification by forward genetics and by using precise reverse-genetics modifications to elucidate their mechanisms. Now, we are devel-oping the capabilities to comprehensively identify and understand

more expansive oligogenic gene-to-phenotype and G 3 E-to-phenotype interactions (seeFigure 1C). Despite these advances, a complete holistic understanding of systems biology is still beset by a number of challenges ahead: (1) there remains a need for higher-quality and more dynamic omics measurements; (2) epistatic gene modeling requires exponentially increasing sample sizes; (3) even the best statistical assurances of causality do not preclude a necessity for experimental validation, at least so far; and (4) there is no ‘‘ideal’’ experimental setup for the study of complex traits. For the first challenge, the next stage is already in sight, with single-cell dynamic profiling (Cai et al., 2006), quan-titative and unbiased metabolite analysis (Fuhrer et al., 2011), larger-scale untargeted SWATH proteomics (Ro¨st et al., 2014), and other such approaches breaking new ground. To the second point, model populations are becoming larger and more refined (Churchill et al., 2004), and G/LOF tools are becoming increas-ingly multiplexed (Chen et al., 2015; Sander and Joung, 2014)— together allowing identification and validation of moderate-scale epistatic interactions. The third obstacle may perhaps be ad-dressed in time, but there remains little indication that accurate mathematical models are on the horizon, except for a few specific biomolecular processes such as protein folding. The last point is more fundamental: a new experiment will always be necessary for testing a new drug or treatment, regardless of the improvements in predictive models or genomics resources.

Finally, further changes in mindset are required. As hypothe-ses become larger and more complex, data and experimental methods must be more readily communicated within and outside of the standard publication cycle. In academic research, this need is now regularly accepted, yet a great deal of data and methods are not made available with publications, either due to privacy concerns (e.g., human data), a lack of a standard database (e.g., raw mass spectrometry data), or even active obfuscation (e.g., in patents, agricultural [Zamir, 2014], or phar-maceutical [Prayle et al., 2012] research). However, given the tremendous progress in resource sharing within the last decade, we expect many of these concerns to diminish in the coming years. There is a revolution in complex trait analysis well under-way, and together the combined applications of systems biology and reductionist approaches are beginning to turn the trickle of genetic discoveries into a steady stream that will usher in a new era for personalized medicine and for environmentally safe genetically modified crops.

AUTHOR CONTRIBUTIONS

E.G.W. and J.A. outlined and designed the manuscript. E.G.W. drafted the manuscript and prepared figures. J.A. approved and edited the manuscript.

ACKNOWLEDGMENTS

(8)

REFERENCES

Albert, F.W., Treusch, S., Shockley, A.H., Bloom, J.S., and Kruglyak, L. (2014). Genetics of single-cell protein abundance variation in large yeast populations. Nature 506, 494–497.

Alonso, J.M., Stepanova, A.N., Leisse, T.J., Kim, C.J., Chen, H., Shinn, P., Ste-venson, D.K., Zimmerman, J., Barajas, P., Cheuk, R., et al. (2003). Genome-wide insertional mutagenesis of Arabidopsis thaliana. Science 301, 653–657.

Andreux, P.A., Williams, E.G., Koutnikova, H., Houtkooper, R.H., Champy, M.F., Henry, H., Schoonjans, K., Williams, R.W., and Auwerx, J. (2012). Sys-tems genetics of metabolism: the use of the BXD murine reference panel for multiscalar integration of traits. Cell 150, 1287–1299.

Arru, L., Francia, E., and Pecchioni, N. (2003). Isolate-specific QTLs of resis-tance to leaf stripe (Pyrenophora graminea) in the ‘Steptoe’ x ‘Morex’ spring barley cross. Theor. Appl. Genet. 106, 668–675.

Auwerx, J., Avner, P., Baldock, R., Ballabio, A., Balling, R., Barbacid, M., Berns, A., Bradley, A., Brown, S., Carmeliet, P., et al. (2004). The European dimension for the mouse genome mutagenesis program. Nat. Genet. 36, 925–927.

Ayyadevara, S., Ayyadevara, R., Hou, S., Thaden, J.J., and Shmookler Reis, R.J. (2001). Genetic mapping of quantitative trait loci governing longevity of Caenorhabditis elegans in recombinant-inbred progeny of a Bergerac-BO x RC301 interstrain cross. Genetics 157, 655–666.

Bailey, D.W. (1981). Recombinant inbred strains and bilineal congenic strains (New York: Academic Press).

Barretina, J., Caponigro, G., Stransky, N., Venkatesan, K., Margolin, A.A., Kim, S., Wilson, C.J., Leha´r, J., Kryukov, G.V., Sonkin, D., et al. (2012). The Cancer Cell Line Encyclopedia enables predictive modelling of anticancer drug sensi-tivity. Nature 483, 603–607.

Barrett, T., Wilhite, S.E., Ledoux, P., Evangelista, C., Kim, I.F., Tomashevsky, M., Marshall, K.A., Phillippy, K.H., Sherman, P.M., Holko, M., et al. (2013). NCBI GEO: archive for functional genomics data sets—update. Nucleic Acids Res. 41, D991–D995.

Bastian, F., Parmentier, G., Roux, J., Moretti, S., Laudet, V., and Robinson-Re-chavi, M. (2008). Bgee: Integrating and comparing heterogeneous transcrip-tome data among species. Lect. N. Bioinformat. 5109, 124–131.

Bennett, B.J., Farber, C.R., Orozco, L., Kang, H.M., Ghazalpour, A., Siemers, N., Neubauer, M., Neuhaus, I., Yordanova, R., Guan, B., et al. (2010). A high-resolution association mapping panel for the dissection of complex traits in mice. Genome Res. 20, 281–290.

Bogardus, C., Baier, L., Permana, P., Prochazka, M., Wolford, J., and Hanson, R. (2002). Identification of susceptibility genes for complex metabolic dis-eases. Ann. N Y Acad. Sci. 967, 1–6.

Burr, B., Burr, F.A., Thompson, K.H., Albertson, M.C., and Stuber, C.W. (1988). Gene mapping with recombinant inbreds in maize. Genetics 118, 519–526.

Cai, L., Friedman, N., and Xie, X.S. (2006). Stochastic protein expression in in-dividual cells at the single molecule level. Nature 440, 358–362.

Capecchi, M.R. (2005). Gene targeting in mice: functional analysis of the mammalian genome for the twenty-first century. Nat. Rev. Genet. 6, 507–512.

Chakravarti, A., Clark, A.G., and Mootha, V.K. (2013). Distilling pathophysi-ology from complex disease genetics. Cell 155, 21–26.

Chen, S., Sanjana, N.E., Zheng, K., Shalem, O., Lee, K., Shi, X., Scott, D.A., Song, J., Pan, J.Q., Weissleder, R., et al. (2015). Genome-wide CRISPR screen in a mouse model of tumor growth and metastasis. Cell 160, 1246–1260.

Chesler, E.J., Lu, L., Wang, J., Williams, R.W., and Manly, K.F. (2004). WebQTL: rapid exploratory analysis of gene expression and genetic networks for brain and behavior. Nat. Neurosci. 7, 485–486.

Churchill, G.A., Airey, D.C., Allayee, H., Angel, J.M., Attie, A.D., Beatty, J., Bea-vis, W.D., Belknap, J.K., Bennett, B., Berrettini, W., et al.; Complex Trait Con-sortium (2004). The Collaborative Cross, a community resource for the genetic analysis of complex traits. Nat. Genet. 36, 1133–1137.

Clark, A.G. (2000). Limits to prediction of phenotypes from knowledge of geno-types. Evol. Biol. 32, 205–224.

Cronin, R.M., Field, J.R., Bradford, Y., Shaffer, C.M., Carroll, R.J., Mosley, J.D., Bastarache, L., Edwards, T.L., Hebbring, S.J., Lin, S., et al. (2014). Phe-nome-wide association studies demonstrating pleiotropy of genetic variants within FTO with and without adjustment for body mass index. Front. Genet.

5, 250.

Cubillos, F.A., Billi, E., Zo¨rgo¨, E., Parts, L., Fargier, P., Omholt, S., Blomberg, A., Warringer, J., Louis, E.J., and Liti, G. (2011). Assessing the complex archi-tecture of polygenic traits in diverged yeast populations. Mol. Ecol. 20, 1401– 1413.

Danhauser, K., Sauer, S.W., Haack, T.B., Wieland, T., Staufner, C., Graf, E., Zschocke, J., Strom, T.M., Traub, T., Okun, J.G., et al. (2012). DHTKD1 muta-tions cause 2-aminoadipic and 2-oxoadipic aciduria. Am. J. Hum. Genet. 91, 1082–1087.

Darvasi, A., and Soller, M. (1995). Advanced intercross lines, an experimental population for fine genetic mapping. Genetics 141, 1199–1207.

Datta, K., Baisakh, N., Ganguly, M., Krishnan, S., Yamaguchi Shinozaki, K., and Datta, S.K. (2012). Overexpression of Arabidopsis and rice stress genes’ inducible transcription factor confers drought and salinity tolerance to rice. Plant Biotechnol. J. 10, 579–586.

De Haan, G., and Van Zant, G. (1999). Genetic analysis of hematopoietic cell cycling in mice suggests its involvement in organismal life span. FASEB J

13, 707–713.

Deeb, S.S., Fajas, L., Nemoto, M., Pihlajama¨ki, J., Mykka¨nen, L., Kuusisto, J., Laakso, M., Fujimoto, W., and Auwerx, J. (1998). A Pro12Ala substitution in PPARgamma2 associated with decreased receptor activity, lower body mass index and improved insulin sensitivity. Nat. Genet. 20, 284–287.

Denny, J.C., Bastarache, L., Ritchie, M.D., Carroll, R.J., Zink, R., Mosley, J.D., Field, J.R., Pulley, J.M., Ramirez, A.H., Bowton, E., et al. (2013). Systematic comparison of phenome-wide association study of electronic medical record data and genome-wide association study data. Nat. Biotechnol. 31, 1102– 1110.

Elmer, G.I., Pieper, J.O., Hamilton, L.R., and Wise, R.A. (2010). Qualitative dif-ferences between C57BL/6J and DBA/2J mice in morphine potentiation of brain stimulation reward and intravenous self-administration. Psychopharma-cology (Berl.) 208, 309–321.

Fenn, J.B., Mann, M., Meng, C.K., Wong, S.F., and Whitehouse, C.M. (1989). Electrospray ionization for mass spectrometry of large biomolecules. Science

246, 64–71.

Flint, J., and Mackay, T.F. (2009). Genetic architecture of quantitative traits in mice, flies, and humans. Genome Res. 19, 723–733.

Flint, J., Valdar, W., Shifman, S., and Mott, R. (2005). Strategies for mapping and cloning quantitative trait genes in rodents. Nat. Rev. Genet. 6, 271–286.

Fuhrer, T., Heer, D., Begemann, B., and Zamboni, N. (2011). High-throughput, accurate mass metabolome profiling of cellular extracts by flow injection-time-of-flight mass spectrometry. Anal. Chem. 83, 7074–7080.

Gelman, R., Watson, A., Bronson, R., and Yunis, E. (1988). Murine chromo-somal regions correlated with longevity. Genetics 118, 693–704.

Goldstein, D.B. (2009). Common genetic variation and human traits. N. Engl. J. Med. 360, 1696–1698.

Harris, T.W., Antoshechkin, I., Bieri, T., Blasiar, D., Chan, J., Chen, W.J., De La Cruz, N., Davis, P., Duesbury, M., Fang, R., et al. (2010). WormBase: a compre-hensive resource for nematode research. Nucleic Acids Res. 38, D463–D467.

Horvath, S. (2013). DNA methylation age of human tissues and cell types. Genome Biol. 14, R115.

Houtkooper, R.H., Mouchiroud, L., Ryu, D., Moullan, N., Katsyuba, E., Knott, G., Williams, R.W., and Auwerx, J. (2013). Mitonuclear protein imbalance as a conserved longevity mechanism. Nature 497, 451–457.

(9)

Jovaisaite, V., Mouchiroud, L., and Auwerx, J. (2014). The mitochondrial unfolded protein response, a conserved stress response pathway with impli-cations in health and disease. J. Exp. Biol. 217, 137–143.

Kamath, R.S., Fraser, A.G., Dong, Y., Poulin, G., Durbin, R., Gotta, M., Kana-pin, A., Le Bot, N., Moreno, S., Sohrmann, M., et al. (2003). Systematic func-tional analysis of the Caenorhabditis elegans genome using RNAi. Nature

421, 231–237.

Kanehisa, M., Goto, S., Sato, Y., Kawashima, M., Furumichi, M., and Tanabe, M. (2014). Data, information, knowledge and principle: back to metabolism in KEGG. Nucleic Acids Res. 42, D199–D205.

Kent, W.J., Sugnet, C.W., Furey, T.S., Roskin, K.M., Pringle, T.H., Zahler, A.M., and Haussler, D. (2002). The human genome browser at UCSC. Genome Res.

12, 996–1006.

Khan, D., Chan, A., Millar, J.L., Girard, I.J., and Belmonte, M.F. (2014). Predict-ing transcriptional circuitry underlyPredict-ing seed coat development. Plant Sci 223, 146–152.

King, E.G., Macdonald, S.J., and Long, A.D. (2012). Properties and power of the Drosophila Synthetic Population Resource for the routine dissection of complex traits. Genetics 191, 935–949.

Kingsmore, S.F., Lindquist, I.E., Mudge, J., and Beavis, W.D. (2007). Genome-wide association studies: progress in identifying genetic biomarkers in com-mon, complex diseases. Biomark. Insights 2, 283–292.

Koutnikova, H., Laakso, M., Lu, L., Combe, R., Paananen, J., Kuulasmaa, T., Kuusisto, J., Ha¨ring, H.U., Hansen, T., Pedersen, O., et al. (2009). Identification of the UBP1 locus as a critical blood pressure determinant using a combination of mouse and human genetics. PLoS Genet. 5, e1000591.

Kover, P.X., Valdar, W., Trakalo, J., Scarcelli, N., Ehrenreich, I.M., Purugganan, M.D., Durrant, C., and Mott, R. (2009). A Multiparent Advanced Generation In-ter-Cross to fine-map quantitative traits in Arabidopsis thaliana. PLoS Genet.

5, e1000551.

Lander, E.S. (2011). Initial impact of the sequencing of the human genome. Na-ture 470, 187–197.

Ledford, H. (2012). Drug bests cystic-fibrosis mutation. Nature 482, 145.

Lee, P.N., Forey, B.A., and Coombs, K.J. (2012). Systematic review with meta-analysis of the epidemiological evidence in the 1900s relating smoking to lung cancer. BMC Cancer 12, 385.

Li, T.K., and Lumeng, L. (1984). Alcohol preference and voluntary alcohol in-takes of inbred rat strains and the National Institutes of Health heterogeneous stock of rats. Alcohol. Clin. Exp. Res. 8, 485–486.

Liu, S.C., Kowalski, S.P., Lan, T.H., Feldmann, K.A., and Paterson, A.H. (1996). Genome-wide high-resolution mapping by recurrent intermating using Arabi-dopsis thaliana as a model. Genetics 142, 247–258.

MacArthur, D.G., Balasubramanian, S., Frankish, A., Huang, N., Morris, J., Walter, K., Jostins, L., Habegger, L., Pickrell, J.K., Montgomery, S.B., et al.; 1000 Genomes Project Consortium (2012). A systematic survey of loss-of-function variants in human protein-coding genes. Science 335, 823–828.

Mackay, T.F., Richards, S., Stone, E.A., Barbadilla, A., Ayroles, J.F., Zhu, D., Casillas, S., Han, Y., Magwire, M.M., Cridland, J.M., et al. (2012). The Drosophila melanogaster Genetic Reference Panel. Nature 482, 173–178.

Marx, V. (2013). Biology: The big challenges of big data. Nature 498, 255–260.

McCarthy, M.I., Abecasis, G.R., Cardon, L.R., Goldstein, D.B., Little, J., Ioan-nidis, J.P., and Hirschhorn, J.N. (2008). Genome-wide association studies for complex traits: consensus, uncertainty and challenges. Nat. Rev. Genet. 9, 356–369.

Nadeau, J.H., Singer, J.B., Matin, A., and Lander, E.S. (2000). Analysing complex genetic traits with chromosome substitution strains. Nat. Genet.

24, 221–225.

Nuzhdin, S.V., Pasyukova, E.G., Dilda, C.L., Zeng, Z.B., and Mackay, T.F. (1997). Sex-specific quantitative trait loci affecting longevity in Drosophila mel-anogaster. Proc. Natl. Acad. Sci. USA 94, 9734–9739.

Ossowski, S., Schneeberger, K., Clark, R.M., Lanz, C., Warthmann, N., and Weigel, D. (2008). Sequencing of natural strains of Arabidopsis thaliana with short reads. Genome Res. 18, 2024–2033.

Park, S.J., Jiang, K., Tal, L., Yichie, Y., Gar, O., Zamir, D., Eshed, Y., and Lipp-man, Z.B. (2014). Optimization of crop productivity in tomato using induced mutations in the florigen pathway. Nat. Genet. 46, 1337–1342.

Peirce, J.L., Lu, L., Gu, J., Silver, L.M., and Williams, R.W. (2004). A new set of BXD recombinant inbred lines from advanced intercross populations in mice. BMC Genet. 5, 7.

Pospisilik, J.A., Schramek, D., Schnidar, H., Cronin, S.J., Nehme, N.T., Zhang, X., Knauf, C., Cani, P.D., Aumayr, K., Todoric, J., et al. (2010). Drosophila genome-wide obesity screen reveals hedgehog as a determinant of brown versus white adipose cell fate. Cell 140, 148–160.

Pravenec, M., Klı´r, P., Kren, V., Zicha, J., and Kunes, J. (1989). An analysis of spontaneous hypertension in spontaneously hypertensive rats by means of new recombinant inbred strains. J. Hypertens. 7, 217–221.

Prayle, A.P., Hurley, M.N., and Smyth, A.R. (2012). Compliance with manda-tory reporting of clinical trial results on ClinicalTrials.gov: cross sectional study. BMJ 344, d7373.

Rockman, M.V., and Kruglyak, L. (2009). Recombinational landscape and pop-ulation genomics of Caenorhabditis elegans. PLoS Genet. 5, e1000419.

Ro¨st, H.L., Rosenberger, G., Navarro, P., Gillet, L., Miladinovic, S.M., Schu-bert, O.T., Wolski, W., Collins, B.C., Malmstro¨m, J., Malmstro¨m, L., and Aebersold, R. (2014). OpenSWATH enables automated, targeted analysis of data-independent acquisition MS data. Nat. Biotechnol. 32, 219–223.

Rual, J.F., Ceron, J., Koreth, J., Hao, T., Nicot, A.S., Hirozane-Kishikawa, T., Vandenhaute, J., Orkin, S.H., Hill, D.E., van den Heuvel, S., and Vidal, M. (2004). Toward improving Caenorhabditis elegans phenome mapping with an ORFeome-based RNAi library. Genome Res. 14 (10B), 2162–2168.

Ryder, E., Ashburner, M., Bautista-Llacer, R., Drummond, J., Webster, J., Johnson, G., Morley, T., Chan, Y.S., Blows, F., Coulson, D., et al. (2007). The DrosDel deletion collection: a Drosophila genomewide chromosomal defi-ciency resource. Genetics 177, 615–629.

Sander, J.D., and Joung, J.K. (2014). CRISPR-Cas systems for editing, regu-lating and targeting genomes. Nat. Biotechnol. 32, 347–355.

Schena, M., Shalon, D., Davis, R.W., and Brown, P.O. (1995). Quantitative monitoring of gene expression patterns with a complementary DNA microar-ray. Science 270, 467–470.

Schilling, R.K., Marschner, P., Shavrukov, Y., Berger, B., Tester, M., Roy, S.J., and Plett, D.C. (2014). Expression of the Arabidopsis vacuolar H+

-pyrophos-phatase gene (AVP1) improves the shoot biomass of transgenic barley and in-creases grain yield in a saline field. Plant Biotechnol. J. 12, 378–386.

Shin, S.Y., Fauman, E.B., Petersen, A.K., Krumsiek, J., Santos, R., Huang, J., Arnold, M., Erte, I., Forgetta, V., Yang, T.P., et al.; Multiple Tissue Human Expression Resource (MuTHER) Consortium (2014). An atlas of genetic influ-ences on human blood metabolites. Nat. Genet. 46, 543–550.

Sivasundar, A., and Hey, J. (2005). Sampling from natural populations with RNAI reveals high outcrossing and population structure in Caenorhabditis el-egans. Curr. Biol. 15, 1598–1602.

Skarnes, W.C., Rosen, B., West, A.P., Koutsourakis, M., Bushell, W., Iyer, V., Mujica, A.O., Thomas, M., Harrow, J., Cox, T., et al. (2011). A conditional knockout resource for the genome-wide study of mouse gene function. Nature

474, 337–342.

Smith, L., and Hood, L. (1987). Mapping and sequencing the human genome: How to proceed. Nat. Biotechnol. 5, 933–939.

Soussi, T., Legros, Y., Lubin, R., Ory, K., and Schlichtholz, B. (1994). Multifac-torial analysis of p53 alteration in human cancer: a review. Int. J. Cancer 57, 1–9.

Subramanian, A., Tamayo, P., Mootha, V.K., Mukherjee, S., Ebert, B.L., Gil-lette, M.A., Paulovich, A., Pomeroy, S.L., Golub, T.R., Lander, E.S., and Me-sirov, J.P. (2005). Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl. Acad. Sci. USA

(10)

Taylor, B.A., Heiniger, H.J., and Meier, H. (1973). Genetic analysis of resistance to cadmium-induced testicular damage in mice. Proc. Soc. Exp. Biol. Med.

143, 629–633.

Valdar, W., Solberg, L.C., Gauguier, D., Burnett, S., Klenerman, P., Cookson, W.O., Taylor, M.S., Rawlins, J.N., Mott, R., and Flint, J. (2006). Genome-wide genetic association of complex traits in heterogeneous stock mice. Nat. Genet. 38, 879–887.

Wang, J., Williams, R.W., and Manly, K.F. (2003). WebQTL: web-based com-plex trait analysis. Neuroinformatics 1, 299–308.

Wasinger, V.C., Cordwell, S.J., Cerpa-Poljak, A., Yan, J.X., Gooley, A.A., Wil-kins, M.R., Duncan, M.W., Harris, R., Williams, K.L., and Humphery-Smith, I. (1995). Progress with gene-product mapping of the Mollicutes: Mycoplasma genitalium. Electrophoresis 16, 1090–1094.

Williams, R.W., Gu, J., Qi, S., and Lu, L. (2001). The genetic structure of recom-binant inbred mice: high-resolution consensus maps for complex trait analysis. Genome Biol 2, RESEARCH0046.

Williams, R.W., Bennett, B., Lu, L., Gu, J., DeFries, J.C., Carosone-Link, P.J., Rikke, B.A., Belknap, J.K., and Johnson, T.E. (2004). Genetic structure of the LXS panel of recombinant inbred mouse strains: a powerful resource for com-plex trait analysis. Mamm. Genome 15, 637–647.

Winzeler, E.A., Shoemaker, D.D., Astromoff, A., Liang, H., Anderson, K., An-dre, B., Bangham, R., Benito, R., Boeke, J.D., Bussey, H., et al. (1999). Func-tional characterization of the S. cerevisiae genome by gene deletion and parallel analysis. Science 285, 901–906.

Wu, Y., Williams, E.G., Dubuis, S., Mottis, A., Jovaisaite, V., Houten, S.M., Arg-mann, C.A., Faridi, P., Wolski, W., Kutalik, Z., et al. (2014). Multilayered ge-netics and omics dissection of mitochondrial activity in a mouse reference population. Cell 158, 1415–1430.

Xu, X.M., and Møller, S.G. (2011). The value of Arabidopsis research in under-standing human disease states. Curr. Opin. Biotechnol. 22, 300–307.

Xu, J., Zhao, Q., Du, P., Xu, C., Wang, B., Feng, Q., Liu, Q., Tang, S., Gu, M., Han, B., and Liang, G. (2010). Developing high throughput genotyped chromo-some segment substitution lines based on population whole-genome re-sequencing in rice (Oryza sativa L.). BMC Genomics 11, 656.

Yamamoto, S., Jaiswal, M., Charng, W.L., Gambin, T., Karaca, E., Mirzaa, G., Wiszniewski, W., Sandoval, H., Haelterman, N.A., Xiong, B., et al. (2014). A drosophila genetic resource of mutants to study mechanisms underlying hu-man genetic diseases. Cell 159, 200–214.

Yeo, G.S., Farooqi, I.S., Aminian, S., Halsall, D.J., Stanhope, R.G., and O’Ra-hilly, S. (1998). A frameshift mutation in MC4R associated with dominantly in-herited human obesity. Nat. Genet. 20, 111–112.

Yuan, J., Njiti, V.N., Meksem, K., Iqbal, M.J., Triwitayakorn, K., Kassem, M.A., Davis, G.T., Schmidt, M.E., and Lightfoot, D.A. (2002). Quantitative trait loci in Two Soybean Recombinant Inbred Line Populations Segregating for Yield and Disease Resistance. Crop Sci. 42, 271–277.

Zamir, D. (2014). Botany. A wake-up call with coffee. Science 345, 1124.

Références

Documents relatifs

Finally, the proposed multi-objective optimization approach take into account several criteria, such as: the minimization of the distance between the simulated nal network state

Integrating QTL controlling fatness, lipid metabolites and gene expressions to genetically dissect the adiposity complex trait in a meat chicken cross. Annual Meeting of the

The other approach to reduce the uncertainty of the output is to optimize the system design by taking directly into account the uncertain parameters probability

We propose various assumptions in the structure of wild species, on the domestication of the edible diploids from hybrids between wild forms, on the direct ancestry of triploids

1 University of Montpellier 1, France; 2 Centre Hospitalier Régional Universitaire, Montpellier; 3 MeDALL (Mechanisms of the Development of Allergy); 4 MACVIA-LR (Contre les

Recombinant antibody libraries have been constructed by cloning antibody heavy- or light- chain variable genes directly from lymphocytes of animals or human and then expressing as

Rather, this fourth profile appeared to be characterized by moderately low levels of global workaholism, moderately high levels of specific working excessively and compulsively,

This research employs a multiple case-study approach to examine the assimilation process at organizations that dealt with complex information systems. Multiple key